Top Banner
applied sciences Article The Magnitude and Waveform of Shock Waves Induced by X-ray Lasers in Water Claudiu Andrei Stan 1,2, * , Koji Motomura 3 , Gabriel Blaj 4 , Yoshiaki Kumagai 3 , Yiwen Li 3 , Daehyun You 3 , Taishi Ono 3 , Armin Kalita 1 , Tadashi Togashi 5 , Shigeki Owada 5 , Kensuke Tono 5 , Makina Yabashi 6 , Tetsuo Katayama 5 and Kiyoshi Ueda 3,6, * 1 Department of Physics, Rutgers University-Newark, Newark, NJ 07102, USA; [email protected] 2 Stanford PULSE Institute, SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA 3 Institute of Multidisciplinary Research for Advanced Materials, Tohoku University, Sendai 980-8577, Japan; [email protected] (K.M.); [email protected] (Y.K.); [email protected] (Y.L.); [email protected] (D.Y.); [email protected] (T.O.) 4 Technology Innovation Directorate, SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA; [email protected] 5 Japan Synchrotron Radiation Research Institute (JASRI), 1-1-1 Kouto, Sayo, Hyogo 679-5198, Japan; [email protected] (T.T.); [email protected] (S.O.); [email protected] (K.T.); [email protected] (T.K.) 6 RIKEN SPring-8 Center, 1-1-1 Kouto, Sayo, Hyogo 679-5148, Japan; [email protected] * Correspondence: [email protected] (C.A.S.); [email protected] (K.U.) Received: 3 February 2020; Accepted: 18 February 2020; Published: 22 February 2020 Abstract: The high energy densities deposited in materials by focused X-ray laser pulses generate shock waves which travel away from the irradiated region, and can generate complex wave patterns or induce phase changes. We determined the time-pressure histories of shocks induced by X-ray laser pulses in liquid water microdrops, by measuring the surface velocity of the microdrops from images recorded during the reflection of the shock at the surface. Measurements were made with ~30 μm diameter droplets using 10 keV X-rays, for X-ray pulse energies that deposited linear energy densities from 3.5 to 120 mJ/m; measurements were also made with ~60 μm diameter drops for a narrower energy range. At a distance of 15 μm from the X-ray beam, the peak shock pressures ranged from 44 to 472 MPa, and the corresponding time-pressure histories of the shocks had a fast quasi-exponential decay with positive pressure durations estimated to range from 2 to 5 ns. Knowledge of the amplitude and waveform of the shock waves enables accurate modeling of shock propagation and experiment designs that either maximize or minimize the eect of shocks. Keywords: X-ray lasers; shock waves; laser ablation 1. Introduction The deposition of energy into materials using pulsed radiation sources has many practical applications including laser surgery [1,2], laser micromachining [35], and chemical synthesis [6]. Besides heating and ablation of the irradiated material, the energy deposited can lead to generation of shock waves if the time scale of energy deposition is shorter than the time in which a sound wave propagates across the irradiated region [1]. These shock waves can carry more than half of the energy deposited by pulsed radiation [7,8] and can lead to unwanted material damage beyond the irradiated region [2]. At the same time, the shock waves produced by pulsed radiation sources are a valuable scientific tool for studying the properties and transformations of materials at high pressures [911] and under large tensions [1217]. Appl. Sci. 2020, 10, 1497; doi:10.3390/app10041497 www.mdpi.com/journal/applsci
18

The Magnitude and Waveform of Shock Waves Induced by X-ray ...

May 25, 2022

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

applied sciences

Article

The Magnitude and Waveform of Shock WavesInduced by X-ray Lasers in Water

Claudiu Andrei Stan 1,2,* , Koji Motomura 3, Gabriel Blaj 4 , Yoshiaki Kumagai 3 , Yiwen Li 3,Daehyun You 3 , Taishi Ono 3, Armin Kalita 1, Tadashi Togashi 5, Shigeki Owada 5,Kensuke Tono 5, Makina Yabashi 6 , Tetsuo Katayama 5 and Kiyoshi Ueda 3,6,*

1 Department of Physics, Rutgers University-Newark, Newark, NJ 07102, USA; [email protected] Stanford PULSE Institute, SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA3 Institute of Multidisciplinary Research for Advanced Materials, Tohoku University, Sendai 980-8577, Japan;

[email protected] (K.M.); [email protected] (Y.K.); [email protected] (Y.L.);[email protected] (D.Y.); [email protected] (T.O.)

4 Technology Innovation Directorate, SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA;[email protected]

5 Japan Synchrotron Radiation Research Institute (JASRI), 1-1-1 Kouto, Sayo, Hyogo 679-5198, Japan;[email protected] (T.T.); [email protected] (S.O.); [email protected] (K.T.);[email protected] (T.K.)

6 RIKEN SPring-8 Center, 1-1-1 Kouto, Sayo, Hyogo 679-5148, Japan; [email protected]* Correspondence: [email protected] (C.A.S.); [email protected] (K.U.)

Received: 3 February 2020; Accepted: 18 February 2020; Published: 22 February 2020�����������������

Abstract: The high energy densities deposited in materials by focused X-ray laser pulses generateshock waves which travel away from the irradiated region, and can generate complex wave patternsor induce phase changes. We determined the time-pressure histories of shocks induced by X-ray laserpulses in liquid water microdrops, by measuring the surface velocity of the microdrops from imagesrecorded during the reflection of the shock at the surface. Measurements were made with ~30 µmdiameter droplets using 10 keV X-rays, for X-ray pulse energies that deposited linear energy densitiesfrom 3.5 to 120 mJ/m; measurements were also made with ~60 µm diameter drops for a narrowerenergy range. At a distance of 15 µm from the X-ray beam, the peak shock pressures ranged from 44to 472 MPa, and the corresponding time-pressure histories of the shocks had a fast quasi-exponentialdecay with positive pressure durations estimated to range from 2 to 5 ns. Knowledge of the amplitudeand waveform of the shock waves enables accurate modeling of shock propagation and experimentdesigns that either maximize or minimize the effect of shocks.

Keywords: X-ray lasers; shock waves; laser ablation

1. Introduction

The deposition of energy into materials using pulsed radiation sources has many practicalapplications including laser surgery [1,2], laser micromachining [3–5], and chemical synthesis [6].Besides heating and ablation of the irradiated material, the energy deposited can lead to generation ofshock waves if the time scale of energy deposition is shorter than the time in which a sound wavepropagates across the irradiated region [1]. These shock waves can carry more than half of the energydeposited by pulsed radiation [7,8] and can lead to unwanted material damage beyond the irradiatedregion [2]. At the same time, the shock waves produced by pulsed radiation sources are a valuablescientific tool for studying the properties and transformations of materials at high pressures [9–11] andunder large tensions [12–17].

Appl. Sci. 2020, 10, 1497; doi:10.3390/app10041497 www.mdpi.com/journal/applsci

Page 2: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 2 of 18

The development of X-ray free-electron laser facilities (XFEL) operating in the keV photon energyregime [18,19] opened new possibilities for depositing large energy densities in materials. Very highintensities (1020 W/cm2) can be achieved with XFEL pulses because X-rays can be focused moretightly than visible light [20], and the femtosecond duration of the pulses allows the creation of highenergy densities, of nonequilibrium states, and the observation of nonlinear X-ray interactions withmatter [21–23].

Given the high intensity and the femtosecond duration of XFEL pulses, they are expected tolaunch shock waves. In samples with thicknesses much smaller than the X-ray absorption length, theshock waves will have a cylindrical symmetry. Such shocks were observed using time-resolved opticalimaging, initially in water microjets [24] and afterwards in liquid microdroplets [17]. The shockslaunched by XFEL pulses may damage soft samples, such as protein crystals carried in liquidjets for serial femtosecond crystallography [25], in experiments performed at XFELs that generateMHz pulse repetition rates. While the initial experiments carried with 1.1 MHz pulse trains at theEuropean XFEL [26] did not observe shock damage in protein crystals [27–29], it remains possible thatshock damage will occur if such experiments are conducted at the maximum possible pulse rate of4.4 MHz [30].

The reflection of XFEL shocks at the surface of liquid jets and drops was observed to lead tospallation in droplets [17] and to the generation of shock trains in jets [24,30]. In both cases thesephenomena had features not observed in similar experiments with pulsed optical lasers, includingthe generation of very large negative pressures in drops [17], and the generation of very large soundintensities in jets [30]. The quantitative characterization of these phenomena had large uncertainties orsignificant noise, which was caused in part by incomplete knowledge of the properties of the initialshock launched by the XFEL pulse, such as the shock waveform.

Here, we report a more precise characterization of shock waves induced by hard XFEL pulses inwater, based on measurements of the surface velocity of water microdrops during the reflection ofthe shock. Improvements in the experimental design allowed us to resolve and measure the decay ofthe shock pressure after the shock discontinuity, while covering a range of one and a half orders ofmagnitude in the XFEL pulse energy.

2. Materials and Methods

X-ray free-electron laser (XFEL) parameters. The experiment was carried out at the SACLA XFEL [19],in the EH4c experimental station on the BL3 beamline [31]. The XFEL pulses had 9.991 ± 0.029 keVphoton energy and an average pulse energy at the XFEL source around 400–500 µJ. The X-ray beamwas focused by a set of Kirkpatrick-Baez (KB) mirrors [32] to a full width half maximum (FWHM) sizeof 1.3 µm horizontally and 1.235 µm vertically.

The experiment also used synchronized femtosecond optical pulses for imaging. The imagingpulses were generated by the SACLA laser system (a Ti:sapphire chirped pulse amplifier system with30 fs duration and 800 nm wavelength) and were delivered to the EH4c experimental station. The lightpulses were frequency doubled to 400 nm to improve the optical resolution and had an energy of 450 µJat the experimental setup.

Experimental setup for measuring the shock wave characteristics. The shock waves were generated bythe focused X-ray laser pulses in spherical water microdrops with 29.9–30.5 µm and 61 µm diameters,and the shock pressures were calculated from the velocity of the droplets’ surface during the reflectionof the shock wave. Figure 1a shows the experimental geometry: The center of the droplets was alignedwith the XFEL beam to better than 1 µm, and multiple droplets were first exposed to XFEL pulses andthen imaged optically using pulsed laser illumination, to determine how the equatorial diameter ofthe drops varied as a function of the time delay. Differentiation of the diameter data with respect tothe time delay provided the velocity of the droplet’s surface in the equatorial region, where the shockwaves were reflected at an angle of incidence close to 90◦, leading to an instantaneous surface velocityequal to twice the instantaneous particle velocity in the shock wave [33]. The surface position as a

Page 3: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 3 of 18

function of the time delays represents the primary data recorded during the experiment, and the shockproperties were calculated from it.Appl. Sci. 2020, 10, x FOR PEER REVIEW 3 of 18

Figure 1. (a) Experimental design. X-ray laser pulses focused to ~1 µm passed through the center of ~30 µm or ~60 µm water microdrops, and launched shock waves. The drops were imaged optically after the X-ray free-electron laser (XFEL) pulse. The photo shows the shock wave in a 30.5-µm diameter drop, 2.29 ns after the arrival of a 41.5 µJ XFEL pulse. To characterize the shock waves, the surface velocity of the drop was measured using multiple droplet images over a range of imaging delays. (b) Experimental setup. In addition to the main imaging system with a 50× objective, a lower resolution system was used for droplet alignment. The pulse energy was measured after the droplets for each pulse.

A top-view schematic of the experimental setup, which was operated at ambient temperature and pressure, is shown in Figure 1b. After the X-ray pulses were reflected by the KB focusing mirrors, they exited the beamline through three beryllium and one Kapton thin windows, and traveled a 215 mm long path in air before being focused. The droplets were overlapped with the beam in the focal plane of the X-rays and were imaged by a high-resolution microscope with a long working distance objective (Mitutoyo M Plan Apo 50×, NA = 0.55). The imaging axis was horizontal and perpendicular to the XFEL beam. An additional low-resolution imaging system was built on the same axis as the XFEL beam, using mirrors with ~3 mm diameter holes for the XFEL beam, and was used in conjunction with the high-resolution microscope to align the droplets to the beam before each run. After passing through the droplet, the energy of the spent XFEL pulse was measured for each pulse by two detectors that collected X-rays scattered from a Kapton foil, and by a third camera-type detector that also absorbed and blocked the beam.

The water droplets were generated by tapered micronozzles (MJ-AT series, MicroFab Inc., Plano, TX, USA) with 15 and 30 µm diameter orifices, equipped with piezoelectric actuators which induced a controlled Rayleigh breakup of the jet into monodisperse droplets. The nozzle control signal was synchronized to the XFEL pulses, to ensure that a droplet was present and properly aligned prior to the arrival of each pulse. The droplet injection was equipped with a temperature control system using circulating liquid from a temperature-controlled bath, which fed heat exchangers for the liquid sample and for a ~1 cm diameter stream of nitrogen gas coaxial with the nozzle. The temperature control system was not enabled because vibrations from the circulating bath significantly increased the jitter of the drop positions, but the gas stream was still used because it reduced the effect of air drafts in the room on the droplet positions. Thus, the temperature of the whole setup stabilized passively to the temperature of the experimental station and was between 26.1 and 26.25 °C. The temperature was measured with two K-type thermocouples, one placed in contact with the nozzle body, and the other in free space, ~1 mm away from the nozzle orifice; the thermocouples reported temperatures within 0.1 °C of each other.

The experimental setup and the data collection procedure are similar to the ones used to investigate cavitation and spallation induced in water by X-ray laser pulses [17], and we refer the

Figure 1. (a) Experimental design. X-ray laser pulses focused to ~1 µm passed through the center of~30 µm or ~60 µm water microdrops, and launched shock waves. The drops were imaged optically afterthe X-ray free-electron laser (XFEL) pulse. The photo shows the shock wave in a 30.5-µm diameter drop,2.29 ns after the arrival of a 41.5 µJ XFEL pulse. To characterize the shock waves, the surface velocity ofthe drop was measured using multiple droplet images over a range of imaging delays. (b) Experimentalsetup. In addition to the main imaging system with a 50× objective, a lower resolution system wasused for droplet alignment. The pulse energy was measured after the droplets for each pulse.

A top-view schematic of the experimental setup, which was operated at ambient temperature andpressure, is shown in Figure 1b. After the X-ray pulses were reflected by the KB focusing mirrors, theyexited the beamline through three beryllium and one Kapton thin windows, and traveled a 215 mmlong path in air before being focused. The droplets were overlapped with the beam in the focal planeof the X-rays and were imaged by a high-resolution microscope with a long working distance objective(Mitutoyo M Plan Apo 50×, NA = 0.55). The imaging axis was horizontal and perpendicular to theXFEL beam. An additional low-resolution imaging system was built on the same axis as the XFELbeam, using mirrors with ~3 mm diameter holes for the XFEL beam, and was used in conjunction withthe high-resolution microscope to align the droplets to the beam before each run. After passing throughthe droplet, the energy of the spent XFEL pulse was measured for each pulse by two detectors thatcollected X-rays scattered from a Kapton foil, and by a third camera-type detector that also absorbedand blocked the beam.

The water droplets were generated by tapered micronozzles (MJ-AT series, MicroFab Inc., Plano,TX, USA) with 15 and 30 µm diameter orifices, equipped with piezoelectric actuators which induceda controlled Rayleigh breakup of the jet into monodisperse droplets. The nozzle control signal wassynchronized to the XFEL pulses, to ensure that a droplet was present and properly aligned prior tothe arrival of each pulse. The droplet injection was equipped with a temperature control system usingcirculating liquid from a temperature-controlled bath, which fed heat exchangers for the liquid sampleand for a ~1 cm diameter stream of nitrogen gas coaxial with the nozzle. The temperature controlsystem was not enabled because vibrations from the circulating bath significantly increased the jitter ofthe drop positions, but the gas stream was still used because it reduced the effect of air drafts in theroom on the droplet positions. Thus, the temperature of the whole setup stabilized passively to thetemperature of the experimental station and was between 26.1 and 26.25 ◦C. The temperature wasmeasured with two K-type thermocouples, one placed in contact with the nozzle body, and the other

Page 4: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 4 of 18

in free space, ~1 mm away from the nozzle orifice; the thermocouples reported temperatures within0.1 ◦C of each other.

The experimental setup and the data collection procedure are similar to the ones used to investigatecavitation and spallation induced in water by X-ray laser pulses [17], and we refer the readers to theprevious work for general information on the experiment. Here, we will focus on the upgrades madeto that setup and to the data collection and analysis.

Modifications made to the setup to improve the quality of data. The experiments were conducted atatmospheric pressure instead of in a vacuum environment. For the purpose of increasing the accuracyof measurements, operation at atmospheric pressure had multiple advantages: (i) The droplets can bealigned better to the beam since vacuum pump vibrations are eliminated; (ii) the droplet temperatureis nearly constant as evaporation, which cools the drops, is much slower at atmospheric pressure thanin vacuum; and (iii) it allows better and faster setup and tuning of the optical imaging and of thedrop injection, and reduces the duration of the adjustments needed during the beamtime. We notethat operation at atmospheric pressure reduces the XFEL intensity through absorption in air, and thatshocks from the drops will be partially transmitted into surrounding gas. Nevertheless, these processesdid not significantly affect our measurements because the transmission of ~10 keV X-rays through0.215 m of air, 88%, was sufficiently large to perform the measurements. Also, the shocks transmittedinto the nitrogen gas stream can be neglected, because the ratio of acoustic impedances of liquidwater and gaseous nitrogen is ~3000, therefore the intensity of the shock transmitted into the gas isapproximately three orders of magnitude smaller than intensity of the incoming shock [34].

Since the magnitude of shocks depends directly on the XFEL pulse energy, the experiment wasdesigned to accurately measure the pulse energy at the sample, and multiple measurements of the pulseenergy were made for each shot. The primary pulse energy measurements were relative, provided bythe two photodiodes (Hamamatsu S3590-09) measuring scattering from the Kapton foil placed afterthe droplets, as shown in Figure 1b. To calibrate the photodiode readings in absolute energy values,a scaling factor was calculated using an absolute energy detector [35] located upstream of the KBfocusing mirrors. The absolute detector was apertured to respond only to photons that could be focusedby the KB mirrors, and thus provided a measurement proportional to the pulse energy at the sample.The proportionality constant was the beam transmission, which was calculated by compounding lossesat the focusing mirror with absorption in air and in the X-ray windows. The absolute energy detectorcould not be used for all data, since in most of the experimental runs, uncalibrated X-ray attenuatorswere inserted after the detector.

The absolute detector and the relative detectors used the same type of photodiodes and the sametype of electronic reading systems [36]. The shot-to-shot measurements of the absolute detector andthe relative detectors were proportional, with a linear correlation coefficient r = 0.9998. We thereforeexpect that (i) the relative accuracy of the pulse energy measurements is close to the one of the absolutedetectors, which was estimated [36] at 10−3, and (ii) the absolute accuracy is also close to the one of theabsolute detectors, which was measured at 3.6% at 9.6 keV by comparison with a primary detectorstandard [37].

To improve the accuracy of measurements of the droplets’ surface position, the numerical aperture(NA) of the main imaging system was reduced from the objective specification (NA = 0.55) to NA = 0.3using irises placed before the condenser lens and in front of the objective lens. The lower NA improvedthe image contrast and made the edge of the drops perfectly dark (i.e., having a pixel signal of 0 in thecamera data), which improved the precision with which the drop surface position was measured.

Data recording and analysis. The data was recorded at an XFEL pulse rate of 30 Hz, in runs thatwere approximately 20 min long. During the runs, the time delay of the illumination laser was scannedat a rate of ~10 ps/s, resulting in probing a time delay range of ~12 ns during a run. The relatively slowscan rate of the time delay resulted in finely spaced time delay data, but it also made the measurementssensitive to slow drifts in the droplet diameter and position.

Page 5: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 5 of 18

The time delay between the XFEL pulse and the illumination of the image could be set staticallywith 1 ps accuracy. During the time delay scans, we recorded this delay for each XFEL shot using atimer-counter (Keysight 53230A, 20 ps specified accuracy). The timer measured the time delay betweenthe XFEL electronic trigger and the signal from a fast photodiode (Hamamatsu G4176-03) that pickedup part of the illumination light. The zero time delay condition between X-rays and illumination wasdetermined before the beamtime using the fast photodiode, and was refined during the beamtime toan accuracy around 2 ps using the onset of the ionization caused by the XFEL beam, at the maximumavailable intensity.

Since the XFEL pulses had a wide distribution of pulse energies, the data was binned by pulseenergy in consecutive logarithmic bins with a 0.05 (5%) relative width, i.e., the data in a bin had pulseenergies ranging from 0.95 to 1 times of the bin maximum. From one run, nine to 11 binned data setswere extracted, according to the condition that each binned set had sufficient data to sample the timedelays with an average time step of ~10 ps.

The drop images had a magnification of 207 nm/pixel, an optical resolution better than 780 nm,and were recorded with a high-speed camera (Phantom Miro M340, Vison Research). The procedurefor determining the droplet diameters is similar to the one described previously [17] with the followingdifferences: (i) We extracted the image data directly from the raw camera data in a compressed 10-bitformat and linearized it, instead of converting the data to 8-bit movie format. (ii) Since the raw dropletimages already had very good edge contrast, it was not necessary to apply a nonlinear rescaling of thepixel data. Images in which the droplet center was misaligned by more than ~0.5 µm from the XFELbeam were removed from further analysis, because for such drops the shock waves arriving at oppositeregions along the drop’s equator had sufficiently different arrival times to affect further analysis.

The operation in open atmosphere and the upgrades to the optical setup and image processingreduced the jitter of drop diameter measurements below 50 nm peak-to-peak in eight out of nine runs.This is a several-fold improvement from the previous study on XFEL-induced spallation [17], wherethe drop diameter data had a jitter ranging from 100 to 250 nm peak-to-peak.

The noise in the drop diameter measurements prevented a point-by-point derivation of the datawith respect to the time delay, as the velocity data would have been extremely noisy. We chose to fitthe diameter data with an exponentially decaying function:

DD(t) − D0(t) = V0 (t − tshock) exp[−(t − tshock)/tdecay], (1)

where t is the time elapsed from the XFEL pulse arrival, DD(t) is the measured diameter of the drop, D0(t)is the diameter of the drop before the shock arrival including a linear correction of the experimentaldrifts in the drop diameter, V0 is the initial diameter expansion velocity upon the arrival of the shockwave, tshock is the arrival time of the shock (measured from the XFEL pulse arrival), and tdecay is thetime in which the surface velocity decays from its peak value to zero. This diameter function leads to aquasi-exponential decay of the surface velocity. It has three fit parameters (V0, tshock, tdecay) that have aclear physical significance and can be quickly estimated from the data to provide starting values forthe fitting. Appendix A contains further details about the fit function and the fitting procedure.

The particle velocity in the shock, up(t), is equal to half of the surface velocity, which is itself halfof the droplet diameter expansion velocity. Thus, the particle velocity is one quarter of the diametervelocity and is given by:

up(t) = (1/4)d/dt[DD(t) − D0(t)], (2)

The initial pressure of the shock, PS(tshock), was derived using the Rankine-Hugoniot conditions [38]and is given by:

PS(tshock) = ρ0us(tshock)up(tshock), (3)

where ρ0 is the water density before the shock arrival, and us is the speed of the shock wave when itarrives at the surface. To calculate the speed of the shock wave we used the empirical relation [39]us(tshock) = c0 + 2up(tshock), where c0 is the speed of sound in water prior to the shock arrival.

Page 6: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 6 of 18

The pressure of the incoming wave after the shock discontinuity, PS(t > tshock), was calculated bynumerically integrating of temporal derivative of the pressure, using the relation:

dPS(t)/dt = ρ0CC[PS(t)]dup(t)/dt, (4)

where CC[PS(t)] is the cylindrical phase velocity [40] in the decaying pressure wave. Equation (4)models the adiabatic evolution of the liquid after the shock discontinuity [40–42] taking into accountthe cylindrical geometry of the shock [43]. Equation (4) and the calculation of the phase velocity arediscussed further in Appendix B.

3. Results

To cover a wide range of pulse energies, multiple runs were recorded with XFEL pulses attenuatedin different amounts by a combination of thin Si wafers and Al foils. Measurements with 29.9–30.5 µmdiameter drops were made for pulse energies ranging from 7.1 µJ to 243 µJ at the droplets. Injection ofthe 61 µm diameter drops was affected by drifts in the undisturbed drop diameters, and these driftswere often larger than the measured signal. Only one experimental run with 61 µm diameter dropshad sufficiently stable drops, and allowed measurements for pulse energies ranging from 47 µJ to 79 µJ.

Raw data sets extracted from one experimental run are shown in Figure 2a. As the pulseenergy increased, (i) the surface displacement became larger, which implies a higher shock pressure,and (ii) shocks arrived at the surface earlier, because the shock velocity increases with the shockamplitude. Figure 2a also shows the slowly-drifting undisturbed drop diameter D0(t), calculatedby linear fitting of the diameter data just before the arrival of the shock. The drift was less thanone thousandth of the drop diameter over the time delays used for measurements (8–10 ns) and thesubtraction of D0(t) from the data was sufficient to correct for the drifts.

Appl. Sci. 2020, 10, x FOR PEER REVIEW 6 of 18

the cylindrical geometry of the shock [43]. Equation (4) and the calculation of the phase velocity are discussed further in Appendix B.

3. Results

To cover a wide range of pulse energies, multiple runs were recorded with XFEL pulses attenuated in different amounts by a combination of thin Si wafers and Al foils. Measurements with 29.9–30.5 µm diameter drops were made for pulse energies ranging from 7.1 µJ to 243 µJ at the droplets. Injection of the 61 µm diameter drops was affected by drifts in the undisturbed drop diameters, and these drifts were often larger than the measured signal. Only one experimental run with 61 µm diameter drops had sufficiently stable drops, and allowed measurements for pulse energies ranging from 47 µJ to 79 µJ.

Raw data sets extracted from one experimental run are shown in Figure 2a. As the pulse energy increased, (i) the surface displacement became larger, which implies a higher shock pressure, and (ii) shocks arrived at the surface earlier, because the shock velocity increases with the shock amplitude. Figure 2a also shows the slowly-drifting undisturbed drop diameter D0(t), calculated by linear fitting of the diameter data just before the arrival of the shock. The drift was less than one thousandth of the drop diameter over the time delays used for measurements (8–10 ns) and the subtraction of D0(t) from the data was sufficient to correct for the drifts.

Figure 2. Experimental measurements and the fit of surface motion data. (a) Typical drop diameter measurements from a single experimental run. For clarity, only four of the 10 extracted data sets are shown. Since these data sets were recorded simultaneously, they exhibit the same drift in the undisturbed droplet diameter, D0(t). (b) The change in drop diameter during the shock reflection, DD(t)−D0(t), for two of the data sets shown in the left panel, and their analytical fits made using Equation (1). To reduce the influence of shock wave reflections, the data is fitted only up to the delay corresponding to the propagation of pressure waves at the speed of sound.

The fitting of data is illustrated in Figure 2b. After subtracting D0(t), the data was fitted in a limited range of time delays, as explained in Appendix A. The fit parameters for all recorded data sets are given in Table A1 in Appendix C, along with the initial shock pressure. Data in Table A1 shows that as the pulse energy increased, the initial shock pressure (PS(tdecay)) increased, the arrival time of the shock (tshock) decreased, and the decay time of the surface velocity (tdecay) increased.

The dependence of the initial shock pressure (the jump pressure) on the pulse energy and on the linear energy density deposited by X-rays is shown in Figure 3 for all data sets. The linear energy deposited is given by the product of the X-ray pulse energy and the X-ray energy-absorption coefficient (493 m−1 at 10 keV for water). At the lowest pulse energies, the relative quality of the fit decreased because the displacement of the interface was approaching the typical jitter and drift of the initial droplet diameter. The reproducibility of the peak pressure measurements, estimated as the

Figure 2. Experimental measurements and the fit of surface motion data. (a) Typical drop diametermeasurements from a single experimental run. For clarity, only four of the 10 extracted data setsare shown. Since these data sets were recorded simultaneously, they exhibit the same drift in theundisturbed droplet diameter, D0(t). (b) The change in drop diameter during the shock reflection,DD(t)−D0(t), for two of the data sets shown in the left panel, and their analytical fits made usingEquation (1). To reduce the influence of shock wave reflections, the data is fitted only up to the delaycorresponding to the propagation of pressure waves at the speed of sound.

The fitting of data is illustrated in Figure 2b. After subtracting D0(t), the data was fitted in alimited range of time delays, as explained in Appendix A. The fit parameters for all recorded data setsare given in Table A1 in Appendix C, along with the initial shock pressure. Data in Table A1 shows

Page 7: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 7 of 18

that as the pulse energy increased, the initial shock pressure (PS(tdecay)) increased, the arrival time ofthe shock (tshock) decreased, and the decay time of the surface velocity (tdecay) increased.

The dependence of the initial shock pressure (the jump pressure) on the pulse energy and on thelinear energy density deposited by X-rays is shown in Figure 3 for all data sets. The linear energydeposited is given by the product of the X-ray pulse energy and the X-ray energy-absorption coefficient(493 m−1 at 10 keV for water). At the lowest pulse energies, the relative quality of the fit decreasedbecause the displacement of the interface was approaching the typical jitter and drift of the initialdroplet diameter. The reproducibility of the peak pressure measurements, estimated as the largestdifference between independent measurements made at the same pulse energies in different runs, wasapproximately 25 MPa.

Appl. Sci. 2020, 10, x FOR PEER REVIEW 7 of 18

largest difference between independent measurements made at the same pulse energies in different runs, was approximately 25 MPa.

Figure 3. Dependence of the peak shock pressure on the pulse energy. At the same pulse energy, the pressure was smaller in the larger drops due to the geometrical decay of the expanding cylindrical shock. The peak pressures in the small drops were fitted to a power-law function of the pulse energy.

For 29.9–30.5 µm diameter drops, the initial shock pressure can be fit with a power-law dependence on the pulse energy Exray, PS(tshock)~Exray0.65. The data for 61 µm diameter drops does not have sufficient dynamic range in the pulse energies for a power law fit.

The temporal evolution of the incoming shock pressure, at the equatorial surface of the drop, was calculated through integration of Equation (4) and is shown in Figure 4 for a subset of measurements. For approximately the same drop diameters, the pressure waveform was similar for different pulse energies and extrapolated to zero pressure ~1 ns after the acoustic time, which is earlier than the time delay when the fitted particle velocity crosses zero (i.e., tshock + tdecay). This behavior is caused by the phase velocity CC[PS(t)] being larger than the speed of sound due to the shock decay and the cylindrical geometry, which following Equation (4) leads to a faster relative decay of the pressure, dPS/PS, than the relative decay of the particle velocity, dup/up.

We note that our calculation of pressure applies to Lagrangian element of fluid located at the surface (i.e., it applies to the same Lagrangian position). Therefore, due to the motion of the surface, the pressure curves in Figure 4 are slightly different from the pressure evolution at the initial position of the surface (the Eulerian position). Also, due to the conservation of energy in cylindrical symmetry, the temporal waveform of the pressure does not have the same shape as the spatial waveform of the pressure.

Figure 3. Dependence of the peak shock pressure on the pulse energy. At the same pulse energy, thepressure was smaller in the larger drops due to the geometrical decay of the expanding cylindricalshock. The peak pressures in the small drops were fitted to a power-law function of the pulse energy.

For 29.9–30.5 µm diameter drops, the initial shock pressure can be fit with a power-law dependenceon the pulse energy Exray, PS(tshock)~Exray

0.65. The data for 61 µm diameter drops does not have sufficientdynamic range in the pulse energies for a power law fit.

The temporal evolution of the incoming shock pressure, at the equatorial surface of the drop, wascalculated through integration of Equation (4) and is shown in Figure 4 for a subset of measurements.For approximately the same drop diameters, the pressure waveform was similar for different pulseenergies and extrapolated to zero pressure ~1 ns after the acoustic time, which is earlier than the timedelay when the fitted particle velocity crosses zero (i.e., tshock + tdecay). This behavior is caused bythe phase velocity CC[PS(t)] being larger than the speed of sound due to the shock decay and thecylindrical geometry, which following Equation (4) leads to a faster relative decay of the pressure,dPS/PS, than the relative decay of the particle velocity, dup/up.

We note that our calculation of pressure applies to Lagrangian element of fluid located at thesurface (i.e., it applies to the same Lagrangian position). Therefore, due to the motion of the surface,the pressure curves in Figure 4 are slightly different from the pressure evolution at the initial positionof the surface (the Eulerian position). Also, due to the conservation of energy in cylindrical symmetry,the temporal waveform of the pressure does not have the same shape as the spatial waveform ofthe pressure.

Page 8: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 8 of 18

Appl. Sci. 2020, 10, x FOR PEER REVIEW 8 of 18

Figure 4. Shock wave waveforms, calculated as described in the text from the fitted surface motion functions. For clarity, the data sets include the shocks with the smallest and largest amplitudes, and a subset of all available waveforms. (a) Waveforms at the surface of small drops, 14.95–15.25 µm from the X-ray beam center. (b) Waveforms at the surface of large drops, 30.5 µm from the X-ray beam.

4. Discussion

The energy carried by the shock wave. The energy carried by the shock can be calculated by integrating the instantaneous wave intensity, IS(t), which is given by:

IS(t) = PS(t)up(t), (5)

Because the shocks launched in the droplets are, to a good approximation, cylindrically symmetric, their energies were expressed as linear energy densities. The linear energy density carried by the shock, from tshock until tacoustic, is shown as a function of the deposited linear energy density in Figure 5a, and the ratio of the shock energy to the deposited energy is shown in Figure 5b. Since the shock wave energies shown in Figure 5 have been integrated only up to tacoustic, they are less than the total energy carried by the shock. Also, the shocks already dissipated a fraction of their energy via irreversible heating of the liquid, when they had large amplitudes shortly after generation [8]. Nevertheless, the trends in the energy carried by shocks can be useful for designing XFEL experiments. Both the shock energy and the fractional amount of energy increased with the XFEL pulse energy, thus the energy of the shock increased faster than linearly with the pulse energy.

Figure 5. The energy carried by the shock waveforms up to tacoustic. (a) Shock energy as a function of deposited energy. (b) Fraction of the absorbed energy carried by the shock waves.

Figure 4. Shock wave waveforms, calculated as described in the text from the fitted surface motionfunctions. For clarity, the data sets include the shocks with the smallest and largest amplitudes, and asubset of all available waveforms. (a) Waveforms at the surface of small drops, 14.95–15.25 µm fromthe X-ray beam center. (b) Waveforms at the surface of large drops, 30.5 µm from the X-ray beam.

4. Discussion

The energy carried by the shock wave. The energy carried by the shock can be calculated by integratingthe instantaneous wave intensity, IS(t), which is given by:

IS(t) = PS(t)up(t), (5)

Because the shocks launched in the droplets are, to a good approximation, cylindrically symmetric,their energies were expressed as linear energy densities. The linear energy density carried by the shock,from tshock until tacoustic, is shown as a function of the deposited linear energy density in Figure 5a,and the ratio of the shock energy to the deposited energy is shown in Figure 5b. Since the shock waveenergies shown in Figure 5 have been integrated only up to tacoustic, they are less than the total energycarried by the shock. Also, the shocks already dissipated a fraction of their energy via irreversibleheating of the liquid, when they had large amplitudes shortly after generation [8]. Nevertheless, thetrends in the energy carried by shocks can be useful for designing XFEL experiments. Both the shockenergy and the fractional amount of energy increased with the XFEL pulse energy, thus the energy ofthe shock increased faster than linearly with the pulse energy.

Appl. Sci. 2020, 10, x FOR PEER REVIEW 8 of 18

Figure 4. Shock wave waveforms, calculated as described in the text from the fitted surface motion functions. For clarity, the data sets include the shocks with the smallest and largest amplitudes, and a subset of all available waveforms. (a) Waveforms at the surface of small drops, 14.95–15.25 µm from the X-ray beam center. (b) Waveforms at the surface of large drops, 30.5 µm from the X-ray beam.

4. Discussion

The energy carried by the shock wave. The energy carried by the shock can be calculated by integrating the instantaneous wave intensity, IS(t), which is given by:

IS(t) = PS(t)up(t), (5)

Because the shocks launched in the droplets are, to a good approximation, cylindrically symmetric, their energies were expressed as linear energy densities. The linear energy density carried by the shock, from tshock until tacoustic, is shown as a function of the deposited linear energy density in Figure 5a, and the ratio of the shock energy to the deposited energy is shown in Figure 5b. Since the shock wave energies shown in Figure 5 have been integrated only up to tacoustic, they are less than the total energy carried by the shock. Also, the shocks already dissipated a fraction of their energy via irreversible heating of the liquid, when they had large amplitudes shortly after generation [8]. Nevertheless, the trends in the energy carried by shocks can be useful for designing XFEL experiments. Both the shock energy and the fractional amount of energy increased with the XFEL pulse energy, thus the energy of the shock increased faster than linearly with the pulse energy.

Figure 5. The energy carried by the shock waveforms up to tacoustic. (a) Shock energy as a function of deposited energy. (b) Fraction of the absorbed energy carried by the shock waves.

Figure 5. The energy carried by the shock waveforms up to tacoustic. (a) Shock energy as a function ofdeposited energy. (b) Fraction of the absorbed energy carried by the shock waves.

Page 9: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 9 of 18

Within the range of parameters of our experiments (PS < 500 MPa), the shock waves were weak,in the sense that the energy lost by increasing the entropy of the shocked water remained small relativeto the total energy of the shock [44]. We thus expect that the energy of the shock decayed slowly as ittraveled between positions corresponding to the radii of the small and large droplets, and Figure 5shows that at the same XFEL pulse energy, the calculated energy was approximately the same for~30 µm and 61 µm diameter drops. The larger energy calculated for the 61 µm drops can be explainedby the experimental uncertainty in energy, which arises from the uncertainty in pressure. We note thatthe maximum conversion ratio of absorbed energy to shock energy shown in Figure 5b, ~12%, falls inbetween the conversion ratios observed for near-field and far-field shock waves generated by pulsedoptical ablation in water [7,8].

The short duration of the shock wave. Two important characteristics of the cylindrical shocksgenerated by XFEL pulses in droplets are their brief duration and the likely decay of pressure tonegative values (tension). The pressures in Figure 4a decayed to less than one tenth of their peakpressure by tacoustic. Since in our experiment the pressure decayed faster than the particle velocity,shocks that led to a maximum in drop diameter, and thus to the particle velocity crossing zero (see theexample in Appendix A), had a negative pressure phase. Extrapolation of the waveforms in Figure 4aindicates that the pressure reached zero between 2 and 5 ns after the shock arrival, depending onthe pulse energy. Multiplying these durations with the speed of sound in water (~1500 m/s) we canestimate the approximate length of the positive waveform (3 to 7.5 µm), which is smaller than the15 µm radius of the drop. Because the positive part of the shock waveform was shorter than the dropradius, its reflections traveled inside the liquid until they no longer overlapped with the positivepressure phase of the incoming shock. Since the shock reflections at a water-air interface are rarefactionwaves [34], pronounced pressure oscillations should occur in jets and drops exposed to XFEL pulses.Two previously observed effects of these pressure oscillations are spallation in droplets [17] and theformation of shock trains in jets [24,30].

Recent numerical simulations of experiments on microscale ablation in water [16,45–47] allow usto compare qualitatively the waveform of shocks produced by XFEL pulses and by optical laser pulses.To simulate the shock dynamics and match it to the experimental conditions, proper initial conditionsneed to be defined. A closed spherical bubble with 8 µm diameter and 6 GPa pressure was usedfor an optical ablation experiment [16,45], and an open cylinder with 0.5 µm diameter and 100 GPapressure was used for an XFEL experiment [47]. In the case of optical ablation with spherical symmetry,the calculated shock waveforms were relatively long and did not reach zero pressure. In contrast,the simulation of the XFEL spallation experiment in drops predicted short waveforms with a negativepressure phase. Short-lived shocks that decay to approximately zero pressure were also predicted insimulations of laser ablation in a ring pattern, in a quasi-2D geometry [46]; in this case the excitedregion was chosen to have a width of 1.6 µm and an initial pressure of 2 GPa, and the simulation didnot include the formation of a gas bubble in the excited region. Overall, these simulations suggest thatthe generation of nanosecond shock waveforms that decay to zero pressure requires that the regionwhere the energy is absorbed is at most a few microns in size, and that the ablation does not lead tohigh-pressure gas trapped inside bubbles.

Cavitation and spallation. Cavitation and spallation were observed in additional experimental runsthat probed time delays longer than the ones we used here for shock wave analysis. Qualitatively,these phenomena were the same as the ones reported previously [17], but quantitative properties suchas the arrival time of the spall wave were different, because they depend on the droplet diameter andon the pulse energy. The present experimental setup should enable measurements of the cavitationpressures with improved accuracy and over a wider range of experimental parameters, because theraw data was less noisy and more accurate than the data previously recorded in vacuum. Our presentwork also indicates the necessity to consider how the short duration of the shock waveform affects thespall dynamics [48]. For example, simulations of XFEL-induced shocks in drops showed an increasein the rise time of the reflected wave as it traveled [47]. An increase in the rise time of the negative

Page 10: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 10 of 18

pressure pulse will lower the strain rate at which the liquid is stretched, and can lower the peak tensionof the negative pressure wave.

Applicability and accuracy of pump-probe optical microscopy for characterization shock waves at themicroscale. The small size of the liquid droplets and their high surface curvature limits the rangeof techniques that can be used to characterize shock wave properties. For example, single-shotinterferometric techniques such as the velocity interferometer system for any reflector (VISAR) [49]cannot be applied to microdrops because interferometry requires a sufficiently flat surface over the areaprobed by the light beam, such that the reflected light has an approximately constant phase. Currently,the minimum beam spot size of commercial interferometers, ~100 µm, is larger than the drops we used.The curvature of the droplet’s surface makes them well suited, however, for methods based on opticalimaging. Even a transparent drop can have a high-contrast dark edge, if the numerical aperture issufficiently small. Spherical surfaces have their maximum lateral extent at a well-defined positionalong the optical axis, and if this position is in focus, a sharp image of the surface can be obtained.

Our results show that a temporal resolution of at least 0.1–1 ns is needed to resolve the shockwaveform in microdrops. This temporal resolution would require a very high-speed framing camerato record the surface velocity in a single shot. Framing cameras were used successfully to resolvethe dynamics of underwater explosions in a millimeter-sized system [50], but for micron-sized dropssuch cameras must be able to operate at GHz frame rates, which is close to the state of the art forcommercial systems. Nevertheless, since liquid micron-sized droplets can be generated reproduciblyat high rates, pump-probe techniques can be used instead. To our knowledge, outside the XFELdroplet ablation experiments, the pump-probe measurement of surface velocities in microscale liquidsystems was applied in a study of cavitation in an optofluidic chip [51], but in that case the surfaceposition was measured only once, 40 ns after the shock arrival, therefore the shock waveforms couldnot be determined.

The accuracy of our shock characterization method is primarily limited by the reproducibilityof the drop diameters. At short time scales, this uncertainty arises from the jitter in drop diametermeasurements, which typically was below 50 nm peak to peak. For a determination of surface velocitywithin a 1 ns interval, this corresponds to a 50 m/s uncertainty in the droplet expansion velocity(or 12.5 m/s for the particle velocity), which in turn corresponds to an uncertainty of approximately20 MPa in the shock pressure. Nevertheless, this type of uncertainty can be reduced by averagingmultiple measurements. At longer time scales, the magnitude and sign of the slow drifts in theunperturbed droplet diameter can become important. In this case the overall accuracy can be estimatedas half the spread of V0 data in Table A1, for measurements made at the same energy but in differentruns; the uncertainty of surface velocity is approximately 25 m/s, one order of magnitude worse thancommercial VISAR instruments.

The cylindrical shock waves produced by hard X-ray laser pulses in liquid water can be detrimentalor beneficial in different types of XFEL studies, and the ability to predict their properties, includingtheir waveforms, is useful in both cases. The brief duration of XFEL-generated shocks could proveuseful for subjecting samples to large pressures that last only ~1 ns, which would allow, for example,to initiate a dynamic transformation in a material but not allow it to complete. Here we providedcomprehensive amplitude and waveform data for shocks propagating in water and passing througha surface 15 µm away from the beam’s center, but the method can be applied to other liquids andpropagation distances. Also, our experimental data could be used as a benchmark for future analyticalor numerical models that might predict the properties of the XFEL-induced shocks for any propagationdistance and any energy of the XFEL pulses.

Author Contributions: C.A.S. conceived the experiment; C.A.S. and K.U. coordinated the experimental team;K.M., T.K., C.A.S., K.T., S.O., T.T. and M.Y. set up the experiment; S.O. and T.T. prepared the femtosecond opticalpulses for imaging liquid droplets; K.M., G.B., Y.K., Y.L., D.Y., T.O., T.K., K.U. and C.A.S. conducted the experiment;C.A.S., G.B., and A.K. analyzed the data; C.A.S. interpreted the data and wrote the paper. All authors have readand agree to the published version of the manuscript.

Page 11: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 11 of 18

Funding: The experiments were supported by the U.S. Department of Energy, Office of Science, Chemical Sciences,Geosciences, and Biosciences Division, by the X-ray Free Electron Laser Priority Strategy Program of the MEXT, bythe Research Program of ‘Dynamic Alliance for Open Innovation Bridging Human, Environment and Materials”in ”Network Joint Research Center for Materials and Devices,’ and by the IMRAM project. Data analysis andinterpretation were primarily supported by startup funds from Rutgers University–Newark.

Acknowledgments: We thank the staff of SACLA for their support and dedicated contributions, and HowardStone for helpful discussions. The XFEL experiments were performed at the BL3 of SACLA with the approval ofthe Japan Synchrotron Radiation Research Institute (JASRI; Proposal No. 2016B8016).

Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design of thestudy; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision topublish the results.

Appendix A

Empirical fitting of the droplet diameter data with an analytical function. Since the experimental dropletdiameter data was recorded from independent single-shot measurements, it included experimentalnoise in the range of 20 to 50 nm peak-to-peak for undisturbed drops. Also, since the time delays weresampled at ~10 ps or less, but the delay data was recorded with a timer-counter with ~20 ps standarddeviation, the delay measurements also had significant shot-to-shot noise. This degree of experimentalnoise precluded a direct differentiation of diameter data to extract surface velocities. Although shockkinematics data can in principle be calculated from the raw data using Savitzki-Golay fitting [17,52],here we chose to fit the raw data using an analytical function because it provided a more compactresult and the fit quality was good across all data.

The fit function, which is given by Equation (1), is the integral of the Friedlander function [53,54]:(1 − z)exp(−z), with z > 0 and z = (t − tshock)/tdecay. The Friedlander function represents a waveformwith a discontinuous pressure jump at z = 0, followed by an immediate decay that lowers the pressureto a minimum that is below the initial pressure. The Friedlander function was found to describeaccurately the pressure-time history from hemispherical blast waves in air [54], and here we implicitlyassumed that this function can also describe with sufficient accuracy the particle velocity-time historyfor the cylindrical shocks in water.

To fit the data with Equation (1), we first subtracted from the raw data the drop diameter prior tothe arrival of the shock, D0(t), which we expressed as linear function of time because it drifted duringthe experimental runs. The fit procedure is illustrated in Figure 2, and Figure A1 shows the sameprocedure for a different experimental run. The run shown in Figure A1 had the largest drift in D0(t)among all measurements presented here. Nevertheless, it was still possible to correct the data for thedrifts because the relative change in the droplet diameter over the fitted time delays was less than0.002, which is smaller than the 0.02 relative spread of undisturbed drop diameters among the runsusing the smaller drops (29.9–30.5 µm diameters).

Variations in the drop diameter between different runs can affect the merging of different data setsbecause in a larger drop the shock will arrive later, and will have a lower pressure, than in a smallerdrop. Nevertheless, for our measurements the drifts and the run-to-run variations in drop diametershad only a small effect on the accuracy of the derived data. For example, Figure 3 shows that the peakpressure scales by less than the inverse of the propagation distance, therefore the relative systematicerror in the peak pressure, caused by a slightly different droplet diameter, will be smaller than therelative differences (~0.02) between droplet diameters in different runs.

Using surface velocity measurements in droplets, the properties of the cylindrical shock generatedby XFEL pulses can be determined directly only up to a limited time delay, because two additionaltypes of pressure waves can arrive at the surface and overlap with the cylindrical shock: (i) Obliqueshock reflections from surface regions away from the equatorial zone, and (ii) if spallation occurs,reflections from the surface of the void generated by spallation. To minimize the chance of measuringthe contribution of additional waves, the data was fitted only up to either (i) the ‘acoustic’ time delaytacoustic, which is equal to the time in which a low-amplitude wave propagates at the speed of sound

Page 12: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 12 of 18

from the center to the equator of an undisturbed drop, or (ii) until the arrival of the spallation reflection,if earlier than the acoustic time. Also, the shock waveforms were computed only up to the acoustic time.Appl. Sci. 2020, 10, x FOR PEER REVIEW 12 of 18

Figure A1. Raw data from the experimental run that exhibited the largest drifts in the initial drop diameter, and the corresponding fits. (a) Drop diameter measurements. For clarity, only four of the 10 extracted data sets are shown. (b) The change in drop diameter during the shock reflection, DD(t)−D0(t), for two of the data sets shown in the left panel.

The diameter data is fitted for time delays that exclude data within a few hundred picoseconds of the shock arrival, see Figures 3b and A1b. In a perfectly aligned drop the slope of the diameter data should change instantaneously upon the shock arrival, but the recorded data shows a gradual increase in slope due to the jitter and the systematic misalignment of the drops. The droplet centers are aligned with the beam to better than 1 µm, but even misalignments of 0.1 µm will make the shocks arrive at measurably different times at the two equatorial surfaces, leading to smearing of the data recorded at delays within a few hundred picoseconds of the shock arrival time in a perfectly centered drop. To choose the first data point for fitting, we first identified the earliest-delay point from where all measurements at longer delays were larger than the peak (diameter-subtracted) diameter of undisturbed drops. These peak values, caused by drop diameter noise, were typically around 20 nm for the ~30-µm drops. For runs collected at the smallest pulse energies, this first point was located after the gradually rising slope and was thus chosen as the first point for fitting. At larger pulse energies, we empirically increased the threshold value up to 100 nm to avoid fitting data near the shock arrival time. The true rise times of the shock waves were too short to be measured. Since the run in which the drops were best aligned to the XFEL beam had an experimental rise time around 75 ps, we estimate that the true rise time was shorter than 75 ps.

Appendix B

Determination of the waveforms of decaying cylindrical shocks from the surface velocity data. Due to the cylindrical symmetry and the pressure decay, a basic determination of the shock pressure from the shock velocity using the Rankine-Hugoniot relations for plane waves may not be accurate [42,55]. For the determination of the initial pressure jump, Equation (3), which is derived for a plane shock, remains appropriate because the shock thickness is much smaller than the curvature of the shock. Using the upper estimate of the shock rise time from experimental data, 75 ps, and a typical shock velocity of 2 km/s for the present data, we estimate a shock thickness smaller than 150 nm, which is less than 0.01 of the shock radius of curvature of ~15 µm at the equatorial surface of a ~30 µm diameter drop. We also note that the assumption of a particle velocity equal to half of the surface velocity [33,56], which is implicit in Equation (2), is likely to be appropriate because the peak shock pressures are below 1 GPa and no phase transitions occur up to this pressure.

The fast decay of the shock requires a specific analysis for the determination of the pressure waveform, and here we chose a method based on Lagrangian wave analysis [41,42] using Fowles and William’s phase velocity concept [40]. The main result is Equation (4), which allows the numerical integration of pressure from its peak value using the particle velocity history [57]. For plane shocks

Figure A1. Raw data from the experimental run that exhibited the largest drifts in the initial dropdiameter, and the corresponding fits. (a) Drop diameter measurements. For clarity, only four of the 10extracted data sets are shown. (b) The change in drop diameter during the shock reflection, DD(t)−D0(t),for two of the data sets shown in the left panel.

The diameter data is fitted for time delays that exclude data within a few hundred picosecondsof the shock arrival, see Figures 3b and A1b. In a perfectly aligned drop the slope of the diameterdata should change instantaneously upon the shock arrival, but the recorded data shows a gradualincrease in slope due to the jitter and the systematic misalignment of the drops. The droplet centersare aligned with the beam to better than 1 µm, but even misalignments of 0.1 µm will make theshocks arrive at measurably different times at the two equatorial surfaces, leading to smearing of thedata recorded at delays within a few hundred picoseconds of the shock arrival time in a perfectlycentered drop. To choose the first data point for fitting, we first identified the earliest-delay point fromwhere all measurements at longer delays were larger than the peak (diameter-subtracted) diameter ofundisturbed drops. These peak values, caused by drop diameter noise, were typically around 20 nmfor the ~30-µm drops. For runs collected at the smallest pulse energies, this first point was located afterthe gradually rising slope and was thus chosen as the first point for fitting. At larger pulse energies,we empirically increased the threshold value up to 100 nm to avoid fitting data near the shock arrivaltime. The true rise times of the shock waves were too short to be measured. Since the run in which thedrops were best aligned to the XFEL beam had an experimental rise time around 75 ps, we estimatethat the true rise time was shorter than 75 ps.

Appendix B

Determination of the waveforms of decaying cylindrical shocks from the surface velocity data. Due tothe cylindrical symmetry and the pressure decay, a basic determination of the shock pressure fromthe shock velocity using the Rankine-Hugoniot relations for plane waves may not be accurate [42,55].For the determination of the initial pressure jump, Equation (3), which is derived for a plane shock,remains appropriate because the shock thickness is much smaller than the curvature of the shock.Using the upper estimate of the shock rise time from experimental data, 75 ps, and a typical shockvelocity of 2 km/s for the present data, we estimate a shock thickness smaller than 150 nm, which isless than 0.01 of the shock radius of curvature of ~15 µm at the equatorial surface of a ~30 µm diameterdrop. We also note that the assumption of a particle velocity equal to half of the surface velocity [33,56],which is implicit in Equation (2), is likely to be appropriate because the peak shock pressures are below1 GPa and no phase transitions occur up to this pressure.

Page 13: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 13 of 18

The fast decay of the shock requires a specific analysis for the determination of the pressurewaveform, and here we chose a method based on Lagrangian wave analysis [41,42] using Fowles andWilliam’s phase velocity concept [40]. The main result is Equation (4), which allows the numericalintegration of pressure from its peak value using the particle velocity history [57]. For plane shocks ina material that is isotropic and does not exhibit phase transitions, the phase velocity CC[PS(t)] is equalto the Lagrangian phase velocity cL[PS(t)], which is equal to the speed of sound in the compressedmaterial multiplied by the ratio of the compressed material density and the zero-pressure density [38].

In a cylindrically symmetric wave, the phase velocity is different from the speed of sound evenfor linear, low-amplitude pressure waves. Basically, this occurs because in a cylindrically symmetricsystem the wave equation has outgoing wave solutions that are not plane waves, but are given by [58]P(r,t) = A H0

(2)(kr)exp(iωt) where A is the wave amplitude and H0(2) is a Hankel function (a linear

combination of Bessel functions). Far from the source (kr� 1), H0(2)(kr) is approximately equal to

(2/π)(kr)−1/2exp(–ikr + π/4) and the phase velocity Cphase is the same as in a harmonic plane wave,Cphase =ω/k. However, near the origin (kr < ~10) the distance between the consecutive roots of thereal part of the Hankel function is larger than π/k, thus the ‘wavelength’ of a cylindrical wave is longerthan 2π/k, and correspondingly near the symmetry axis Cphase is larger thanω/k.

We derived the phase velocity to be used in Equation (4), CC[PS(t)], by imposing that (i) both massand momentum were conserved in the cylindrical wave and (ii) the pressure and density were relatedby an adiabatic equation of state.

For the equation of state of water, we used a modified Tait equation of state (see the SupplementalMaterial of Reference [16]):

(P + P∞)/(P0 + P∞) = (ρ/ρ0)γ, (A1)

where P and ρ are the absolute pressure and the density, P0 = 1.01325 × 105 Pa and ρ0 = 996.7 kg/m3 arethe pressure and density before the shock (i.e., water at atmospheric pressure and 26.2 ◦C). Equation (A1)is referenced to the liquid density ρ0 at atmospheric pressure, which makes it more convenient to usethan the more common form of the Tait equation [59], which is referenced to the liquid density at zeropressure. We used a value γ = 7.15, as employed in other studies [16]. The value of the constant P∞, P∞= 3.1258 × 108 Pa, was determined by requiring that the speed of sound c0 at 26.2 ◦C, c0 = 1497.7 m/s,is equal to the speed of sound ceos calculated from the equation of state (Equation (A1)):

ceos2 = (∂P/∂ρ)S = γ(P + P∞)/ρ, (A2)

Following Fowles’ derivation of the conservation relations for stress waves in sphericalcoordinates [43], we derived the conservation laws for density and pressure in cylindrical coordinates:

dρ = (ρ2r)/(ρ0r0Cu) du + (ρu)/r dt, (A3)

dP = (ρ0r0CP)/r du, (A4)

where ρ, P, u and r are the density, pressure, particle velocity, and radial position of the Lagrangian fluidelement, ρ0, P0, and r0 the corresponding values before the shock arrival. The two phase velocities [40]are defined by Cu = (∂h/∂t)u and CP = (∂h/∂t)P, where h is the Lagrangian coordinate of the fluidelement. Since the flow of the liquid after the shock discontinuity is adiabatic, and since no phasetransitions occur during flow, we assumed that the phase velocities are equal: Cu = CP = CC(P).

At any given pressure P, Equation (A2) can be used to derive the density variation dρeos = dP/ceos2

corresponding to the differential change in pressure dP given by Equation (A4). The condition thatthe change in density should be the same when calculated by Equation (A3) as when calculated fromEquation (A4) via the equation of state (i.e., dρ = dρeos) leads to a quadratic equation in CC(P), fromwhich we calculated the phase velocity.

Finally, the shock waveforms PS(t) were calculated by numerical integration of Equation (4) fromtshock until tacoustic. The initial pressure was calculated using Equation (3). The analytical fit of the

Page 14: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 14 of 18

surface position, see Equations (1) and (2), was used to determine the particle position r, the particlevelocity u, and its derivative du/dt. The phase velocity CC(PS(t)) was calculated at each time step usingthe computed values of r, u, and du/dt at the same time step, and the pressure PS at the previous timestep. The numerical integration was done in 1000 time steps using a computer code written in Matlab(MATLAB R2018a, The MathWorks, Inc., Natick, MA, USA).

Appendix C

The fitting parameters for the particle velocity data for all measurements are given in Table A1,along with the initial shock wave pressures. For droplets with 29.9–30.5µm diameters, 79 measurementswere done for 55 distinct XFEL pulse energies. For droplets with 60.5 µm diameters, 11 measurementswere done for 11 distinct XFEL pulse energies. The standard deviations given in Table A1 are thestandard deviations of the fit parameters as reported by the software used for fitting (Igor Pro 8,Wavemetrics Inc.) and do not account for systematic errors, such as second-order effects of the drifts inthe undisturbed drop diameter.

Table A1. Measurement conditions, fit parameters for particle velocities (see Equation (1)), and theinitial shock wave pressures PS(tshock). Ddrop is the undisturbed drop diameter just before the arrivalof the shock wave, and Exray is the energy of the X-ray pulse arriving at the droplets. The standarddeviations represent the statistical uncertainties of the fit parameters.

Ddrop [µm] Exray [µJ] V0 [m/s] σv0 [m/s] tshock [ns] σts [ns] tdecay [ns] σtd [ns] PS(tshock) [MPa]

30.2 7.1 113 16 8.991 0.044 1.82 0.43 4430.2 7.4 123 15 8.943 0.039 1.72 0.34 4830.2 7.8 119 12 8.883 0.034 2.03 0.36 4630.2 8.2 121 15 8.839 0.047 2.11 0.43 4730.2 8.7 131 12 8.819 0.033 2.05 0.31 5130.2 9.1 127 10 8.765 0.032 2.55 0.39 5030.2 9.6 142 11 8.732 0.033 2.27 0.29 5630.2 10.1 139 9 8.677 0.028 2.61 0.32 5430.2 10.7 142 7 8.638 0.019 2.81 0.31 5630.2 11.2 145 11 8.592 0.030 2.97 0.50 5730.2 15.3 230 7 8.242 0.012 2.09 0.08 9230.5 15.3 193 7 8.339 0.016 2.66 0.15 7730.5 16.1 202 6 8.289 0.015 2.75 0.14 8130.2 16.1 212 6 8.139 0.011 2.72 0.12 8530.5 16.1 201 7 8.321 0.015 2.56 0.14 8030.5 16.9 207 5 8.225 0.011 2.88 0.12 8330.5 16.9 203 5 8.265 0.012 2.73 0.11 8130.2 16.9 206 5 8.069 0.010 3.14 0.12 8230.5 17.8 208 6 8.203 0.014 2.86 0.13 8330.2 17.8 232 5 8.061 0.011 2.85 0.10 9330.5 17.8 218 5 8.178 0.011 2.93 0.11 8730.5 18.7 221 4 8.102 0.010 3.17 0.11 8830.2 18.7 235 4 7.981 0.009 3.06 0.09 9530.5 18.7 222 4 8.163 0.008 2.88 0.09 8930.2 19.7 238 4 7.910 0.008 3.34 0.09 9630.5 19.7 234 5 8.071 0.011 3.30 0.12 9430.5 19.7 229 4 8.124 0.008 3.02 0.09 9230.5 20.8 254 4 8.026 0.007 3.18 0.08 10330.5 20.8 248 4 8.092 0.008 2.94 0.07 10030.2 20.8 237 4 7.839 0.008 3.76 0.11 9530.2 21.8 242 4 7.755 0.009 4.07 0.14 9830.5 21.8 250 4 7.947 0.009 3.63 0.11 10130.5 21.8 255 4 8.037 0.007 3.11 0.07 10330.2 23.0 266 5 7.749 0.010 3.92 0.13 10830.5 23.0 264 4 7.981 0.007 3.27 0.08 10730.5 23.0 264 5 7.907 0.010 3.68 0.13 10730.2 24.2 273 7 7.690 0.013 4.18 0.20 11130.5 24.2 292 6 7.870 0.012 3.44 0.12 12030.5 24.2 240 8 7.919 0.018 3.43 0.20 97

Page 15: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 15 of 18

Table A1. Cont.

Ddrop [µm] Exray [µJ] V0 [m/s] σv0 [m/s] tshock [ns] σts [ns] tdecay [ns] σtd [ns] PS(tshock) [MPa]

30.5 24.2 272 4 7.929 0.008 3.43 0.09 11130.5 25.5 234 5 7.841 0.011 4.16 0.19 9430.5 25.5 288 5 7.883 0.009 3.45 0.10 11830.5 26.8 260 5 7.811 0.013 3.80 0.14 10530.5 28.2 279 5 7.762 0.009 3.78 0.11 11430.5 29.7 287 4 7.699 0.008 3.96 0.10 11730.5 31.3 298 4 7.652 0.007 4.14 0.10 12230.5 32.9 308 3 7.591 0.006 4.33 0.08 12730.5 34.7 313 4 7.527 0.007 4.67 0.11 12930.5 36.5 331 4 7.486 0.008 4.71 0.12 13730.5 38.4 344 5 7.420 0.011 4.81 0.14 14330.0 52.2 451 5 6.936 0.008 5.20 0.12 19430.0 55.0 457 4 6.870 0.006 5.56 0.09 19630.0 57.9 471 4 6.812 0.006 5.63 0.09 20330.0 60.9 486 3 6.751 0.005 5.67 0.08 21130.0 64.1 500 3 6.686 0.004 5.74 0.06 21830.0 67.5 514 3 6.630 0.004 5.92 0.06 22530.0 71.1 531 3 6.581 0.004 6.01 0.06 23330.0 74.8 554 3 6.539 0.004 5.95 0.06 24530.0 78.7 566 3 6.487 0.004 6.11 0.07 25130.1 107.1 663 7 6.203 0.080 7.03 0.16 30230.1 112.8 673 6 6.142 0.007 7.36 0.15 30830.1 118.7 674 5 6.080 0.006 7.93 0.15 30830.1 124.9 688 5 6.032 0.006 8.18 0.16 31630.1 131.5 700 4 5.981 0.004 8.33 0.14 32330.1 138.4 733 5 5.942 0.004 7.78 0.13 34130.1 145.7 754 5 5.898 0.004 7.67 0.12 35228.9 153.4 815 8 5.823 0.006 7.19 0.17 38730.1 153.4 778 4 5.859 0.004 7.54 0.11 36630.1 161.5 781 5 5.796 0.005 7.90 0.13 36828.9 161.5 804 7 5.757 0.006 8.09 0.17 38130.1 170.0 797 6 5.753 0.006 8.02 0.17 37728.9 170.0 825 6 5.712 0.005 8.06 0.17 39328.9 178.9 851 5 5.669 0.004 7.94 0.14 40828.9 188.3 864 6 5.613 0.004 8.21 0.16 41528.9 198.2 887 5 5.577 0.004 8.26 0.14 42928.9 208.7 897 5 5.520 0.004 8.48 0.16 43529.9 219.7 916 6 5.473 0.004 8.59 0.18 44729.9 231.2 943 7 5.429 0.004 8.58 0.20 46329.9 243.4 957 10 5.377 0.006 8.75 0.30 47261.0 47.1 238 4 16.507 0.031 10.58 0.27 9661.0 49.6 252 3 16.464 0.025 10.43 0.21 10261.0 52.2 257 3 16.369 0.022 10.92 0.20 10461.0 55.0 266 3 16.295 0.021 11.09 0.18 10861.0 57.9 274 2 16.172 0.017 11.40 0.16 11161.0 60.9 279 2 16.064 0.018 11.68 0.17 11461.0 64.1 291 2 15.997 0.017 11.57 0.15 11961.0 67.5 305 2 15.929 0.014 11.33 0.12 12561.0 71.1 317 2 15.860 0.014 11.18 0.12 13161.0 74.8 332 2 15.801 0.014 10.91 0.12 13861.0 78.7 344 3 15.712 0.016 10.76 0.14 143

References

1. Vogel, A.; Venugopalan, V. Mechanisms of Pulsed Laser Ablation of Biological Tissues. Chem. Rev. 2003,103, 577–644. [CrossRef]

2. Vogel, A.; Noack, J.; Huttman, G.; Paltauf, G. Mechanisms of femtosecond laser nanosurgery of cells andtissues. Appl. Phys. B 2005, 81, 1015–1047. [CrossRef]

3. Chichkov, B.N.; Momma, C.; Nolte, S.; von Alvensleben, F.; Tunnermann, A. Femtosecond, picosecond andnanosecond laser ablation of solids. Appl. Phys. A 1996, 63, 109–115. [CrossRef]

4. Nolte, S.; Momma, C.; Jacobs, H.; Tunnermann, A.; Chichkov, B.N.; Wellegehausen, B.; Welling, H. Ablationof metals by ultrashort laser pulses. J. Opt. Soc. Am. B Opt. Phys. 1997, 14, 2716–2722. [CrossRef]

Page 16: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 16 of 18

5. Liu, X.; Du, D.; Mourou, G. Laser ablation and micromachining with ultrashort laser pulses. IEEE J. QuantumElectron. 1997, 33, 1706–1716. [CrossRef]

6. Yang, G.W. Laser ablation in liquids: Applications in the synthesis of nanocrystals. Prog. Mater. Sci. 2007,52, 648–698. [CrossRef]

7. Vogel, A.; Lauterborn, W. Acoustic Transient Generation by Laser-Produced Cavitation Bubbles near SolidBoundaries. J. Acoust. Soc. Am. 1988, 84, 719–731. [CrossRef]

8. Vogel, A.; Noack, J. Shock-wave energy and acoustic energy dissipation after laser-induced breakdown.In Laser-Tissue Interaction IX; International Society for Optics and Photonics: Bellingham, WA, USA, 1998;pp. 180–189.

9. Anderhol, N.C. Laser-generated stress waves. Appl. Phys. Lett. 1970, 16, 113–115. [CrossRef]10. Smith, R.F.; Eggert, J.H.; Jeanloz, R.; Duffy, T.S.; Braun, D.G.; Patterson, J.R.; Rudd, R.E.; Biener, J.; Lazicki, A.E.;

Hamza, A.V.; et al. Ramp compression of diamond to five terapascals. Nature 2014, 511, 330–333. [CrossRef]11. Meyers, M.A.; Gregori, F.; Kad, B.K.; Schneider, M.S.; Kalantar, D.H.; Remington, B.A.; Ravichandran, G.;

Boehly, T.; Wark, J.S. Laser-induced shock compression of monocrystalline copper: Characterization andanalysis. Acta Mater. 2003, 51, 1211–1228. [CrossRef]

12. Carlson, G.A. Dynamic tensile strength of mercury. J. Appl. Phys. 1975, 46, 4069–4070. [CrossRef]13. Fortov, V.E.; Kostin, V.V.; Eliezer, S. Spallation of metals under laser irradiation. J. Appl. Phys. 1991,

70, 4524–4531. [CrossRef]14. Tamura, H.; Kohama, T.; Kondo, K.; Yoshida, M. Femtosecond-laser-induced spallation in aluminum. J. Appl.

Phys. 2001, 89, 3520–3522. [CrossRef]15. Paltauf, G.; Schmidt-Kloiber, H. Microcavity dynamics during laser-induced spallation of liquids and gels.

Appl. Phys. A 1996, 62, 303–311. [CrossRef]16. Ando, K.; Liu, A.Q.; Ohl, C.D. Homogeneous Nucleation in Water in Microfluidic Channels. Phys. Rev. Lett.

2012, 109, 044501. [CrossRef]17. Stan, C.A.; Willmott, P.R.; Stone, H.A.; Koglin, J.E.; Liang, M.; Aquila, A.L.; Robinson, J.S.; Gumerlock, K.L.;

Blaj, G.; Sierra, R.G.; et al. Negative Pressures and Spallation in Water Drops Subjected to Nanosecond ShockWaves. J. Phys. Chem. Lett. 2016, 7, 2055–2062. [CrossRef]

18. Emma, P.; Akre, R.; Arthur, J.; Bionta, R.; Bostedt, C.; Bozek, J.; Brachmann, A.; Bucksbaum, P.; Coffee, R.;Decker, F.J.; et al. First lasing and operation of an ångstrom-wavelength free-electron laser. Nat. Photon.2010, 4, 641–647. [CrossRef]

19. Ishikawa, T.; Aoyagi, H.; Asaka, T.; Asano, Y.; Azumi, N.; Bizen, T.; Ego, H.; Fukami, K.; Fukui, T.;Furukawa, Y.; et al. A compact X-ray free-electron laser emitting in the sub-ångström region. Nat. Photon.2012, 6, 540–544. [CrossRef]

20. Mimura, H.; Yumoto, H.; Matsuyama, S.; Koyama, T.; Tono, K.; Inubushi, Y.; Togashi, T.; Sato, T.; Kim, J.;Fukui, R.; et al. Generation of 10ˆ20 Wcmˆ-2 hard X-ray laser pulses with two-stage reflective focusingsystem. Nat. Commun. 2014, 5, 3539. [CrossRef]

21. Vinko, S.M.; Ciricosta, O.; Cho, B.I.; Engelhorn, K.; Chung, H.K.; Brown, C.R.D.; Burian, T.; Chalupsky, J.;Falcone, R.W.; Graves, C.; et al. Creation and diagnosis of a solid-density plasma with an X-ray free-electronlaser. Nature 2012, 482, 59–62. [CrossRef]

22. Yoneda, H.; Inubushi, Y.; Yabashi, M.; Katayama, T.; Ishikawa, T.; Ohashi, H.; Yumoto, H.; Yamauchi, K.;Mimura, H.; Kitamura, H. Saturable absorption of intense hard X-rays in iron. Nat. Commun. 2014, 5, 5080.[CrossRef] [PubMed]

23. Beyerlein, K.R.; Jonsson, H.O.; Alonso-Mori, R.; Aquila, A.; Barty, S.; Barty, A.; Bean, R.; Koglin, J.E.;Messerschmidt, M.; Ragazzon, D.; et al. Ultrafast nonthermal heating of water initiated by an X-rayFree-Electron Laser. Proc. Natl. Acad. Sci. USA 2018, 115, 5652–5657. [CrossRef] [PubMed]

24. Stan, C.A.; Milathianaki, D.; Laksmono, H.; Sierra, R.G.; McQueen, T.A.; Messerschmidt, M.; Williams, G.J.;Koglin, J.E.; Lane, T.J.; Hayes, M.J.; et al. Liquid explosions induced by X-ray laser pulses. Nat. Phys. 2016,12, 966–971. [CrossRef]

25. Chapman, H.N.; Fromme, P.; Barty, A.; White, T.A.; Kirian, R.A.; Aquila, A.; Hunter, M.S.; Schulz, J.;DePonte, D.P.; Weierstall, U.; et al. Femtosecond X-ray protein nanocrystallography. Nature 2011, 470, 73–77.[CrossRef]

26. Altarelli, M. The European X-ray Free-Electron Laser: Toward an ultra-bright, high repetition-rate x-raysource. High Power Laser Sci. 2015, 3, e18. [CrossRef]

Page 17: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 17 of 18

27. Grünbein, M.L.; Bielecki, J.; Gorel, A.; Stricker, M.; Bean, R.; Cammarata, M.; Doerner, K.; Fröhlich, L.;Hartmann, L.; Hauf, S.; et al. Megahertz Data Collection from Protein Microcrystals at an X-ray Free-ElectronLaser. Nat. Commun. 2018, 9, 3847. [CrossRef]

28. Wiedorn, M.O.; Oberthür, D.; Bean, R.; Schubert, R.; Werner, N.; Abbey, B.; Aepfelbacher, M.; Adriano, A.;Allahgholi, A.; Al-Qudami, N.; et al. Megahertz serial crystallography. Nat. Commun. 2018, 9, 4025.[CrossRef]

29. Yefanov, O.; Oberthür, D.; Bean, R.; Wiedorn, M.O.; Knoska, J.; Pena, G.; Awel, S.; Gumprecht, L.;Domaracky, M.; Sarrou, I.; et al. Evaluation of serial crystallographic structure determination withinmegahertz pulse trains. Struct. Dyn. 2019, 6, 064702. [CrossRef]

30. Blaj, G.; Liang, M.N.; Aquila, A.L.; Willmott, P.R.; Koglin, J.E.; Sierra, R.G.; Robinson, J.S.; Boutet, S.; Stan, C.A.Generation of high-intensity ultrasound through shock propagation in liquid jets. Phys. Rev. Fluids 2019,4, 043401. [CrossRef]

31. Tono, K.; Togashi, T.; Inubushi, Y.; Sato, T.; Katayama, T.; Ogawa, K.; Ohashi, H.; Kimura, H.; Takahashi, S.;Takeshita, K.; et al. Beamline, experimental stations and photon beam diagnostics for the hard x-ray freeelectron laser of SACLA. New J. Phys. 2013, 15, 083035. [CrossRef]

32. Yumoto, H.; Mimura, H.; Koyama, T.; Matsuyama, S.; Tono, K.; Togashi, T.; Inubushi, Y.; Sato, T.; Tanaka, T.;Kimura, T.; et al. Focusing of X-ray free-electron laser pulses with reflective optics. Nat. Photon. 2013,7, 43–47. [CrossRef]

33. Walsh, J.M.; Christian, R.H. Equation of State of Metals from Shock Wave Measurements. Phys. Rev. 1955,97, 1544–1556. [CrossRef]

34. Henderson, L.F. On the refraction of shock waves. J. Fluid Mech. 1989, 198, 365–386. [CrossRef]35. Tono, K.; Kudo, T.; Yabashi, M.; Tachibana, T.; Feng, Y.P.; Fritz, D.; Hastings, J.; Ishikawa, T. Single-shot

beam-position monitor for x-ray free electron laser. Rev. Sci. Instrum. 2011, 82, 023108. [CrossRef] [PubMed]36. Kudo, T.; Tono, K.; Yabashi, M.; Togashi, T.; Sato, T.; Inubushi, Y.; Omodani, M.; Kirihara, Y.; Matsushita, T.;

Kobayashi, K.; et al. A photodiode amplifier system for pulse-by-pulse intensity measurement of an x-rayfree electron laser. Rev. Sci. Instrum. 2012, 83, 043108. [CrossRef] [PubMed]

37. Kato, M.; Tanaka, T.; Kurosawa, T.; Saito, N.; Richter, M.; Sorokin, A.A.; Tiedtke, K.; Kudo, T.; Tono, K.;Yabashi, M.; et al. Pulse energy measurement at the hard x-ray laser in Japan. Appl. Phys. Lett. 2012,101, 023503. [CrossRef]

38. Grady, D. Shocks and structured waves. In Physics of Shock and Impact; IOP Publishing Ltd.: Bristol, UK, 2017.39. Nagayama, K.; Mori, Y.; Shimada, K.; Nakahara, M. Shock Hugoniot compression curve for water up to 1

GPa by using a compressed gas gun. J. Appl. Phys. 2002, 91, 476–482. [CrossRef]40. Fowles, R.; Williams, R.F. Plane Stress Wave Propagation in Solids. J. Appl. Phys. 1970, 41, 360–363. [CrossRef]41. Aidun, J.B.; Gupta, Y.M. Analysis of Lagrangian Gauge Measurements of Simple and Nonsimple Plane-Waves.

J. Appl. Phys. 1991, 69, 6998–7014. [CrossRef]42. Grady, D. Shock wave analysis and applications. In Physics of Shock and Impact; IOP Publishing Ltd.: Bristol,

UK, 2017.43. Fowles, R. Conservation Relations for Spherical and Cylindrical Stress Waves. J. Appl. Phys. 1970,

41, 2740–2741. [CrossRef]44. Rice, M.H.; Walsh, J.M. Equation of State of Water to 250 Kilobars. J. Chem. Phys. 1957, 26, 824–830. [CrossRef]45. Lyu, X.X.; Pan, S.C.; Hu, X.Y.; Adams, N.A. Numerical investigation of homogeneous cavitation nucleation

in a microchannel. Phys. Rev. Fluids 2018, 3, 064303. [CrossRef]46. Veysset, D.; Gutierrez-Hernandez, U.; Dresselhaus-Cooper, L.; De Colle, F.; Kooi, S.; Nelson, K.A.;

Quinto-Su, P.A.; Pezeril, T. Single-bubble and multibubble cavitation in water triggered by laser-drivenfocusing shock waves. Phys. Rev. E 2018, 97, 053112. [CrossRef] [PubMed]

47. Paula, T.; Adami, S.; Adams, N.A. Analysis of the early stages of liquid-water-drop explosion by numericalsimulation. Phys. Rev. Fluids 2019, 4, 044003. [CrossRef]

48. Cai, Y.; Wu, H.A.; Luo, S.N. Spall strength of liquid copper and accuracy of the acoustic method. J. Appl.Phys. 2017, 121, 105901. [CrossRef]

49. Barker, L.; Hollenbach, R. Laser interferometer for measuring high velocities of any reflecting surface. J. Appl.Phys. 1972, 43, 4669–4675. [CrossRef]

50. Menezes, V.; Hosseini, H.; Moosavi-Nejad, S.; Irimpan, K.J.; Akiyama, H. Motion of free-surface ofshock-compressed water on emergence of rarefaction. Appl. Phys. Lett. 2015, 107, 143701. [CrossRef]

Page 18: The Magnitude and Waveform of Shock Waves Induced by X-ray ...

Appl. Sci. 2020, 10, 1497 18 of 18

51. Li, Z.G.; Xiong, S.; Chin, L.K.; Ando, K.; Zhang, J.B.; Liu, A.Q. Water’s tensile strength measured using anoptofluidic chip. Lab Chip 2015, 15, 2158–2161. [CrossRef]

52. Savitzky, A.; Golay, M.J.E. Smoothing and Differentiation of Data by Simplified Least Squares Procedures.Anal. Chem. 1964, 36, 1627–1639. [CrossRef]

53. Friedlander, F.G. The Diffraction of Sound Pulses 1. Diffraction by a Semi-Infinite Plane. Proc. R. Soc. Lond.Ser. A 1946, 186, 322–344.

54. Dewey, J.M. The Friedlander Equations. In Blast Effects: Physical Properties of Shock Waves, 1st ed.; Sochet, I.,Ed.; Springer: Cham, Switzerland, 2018; pp. 37–55.

55. Krehl, P.O.K. The classical Rankine-Hugoniot jump conditions, an important cornerstone of modern shockwave physics: Ideal assumptions vs. reality. Eur. Phys. J. H 2015, 40, 159–204. [CrossRef]

56. Luo, S.N.; Han, L.B.; Xie, Y.; An, Q.; Zheng, L.Q.; Xia, K.W. The relation between shock-state particle velocityand free surface velocity: A molecular dynamics study on single crystal Cu and silica glass. J. Appl. Phys.2008, 103, 093530. [CrossRef]

57. Rothman, S.D.; Davis, J.P.; Maw, J.; Robinson, C.M.; Parker, K.; Palmer, J. Measurement of the principalisentropes of lead and lead-antimony alloy to similar to 400 kbar by quasi-isentropic compression. J. Phys. DAppl. Phys. 2005, 38, 733–740. [CrossRef]

58. Kinsler, L.E.; Frey, A.R.; Coppens, A.B.; Sanders, J.V. The acoustic wave equation and simple solutions.In Fundamentals of Acoustics, 4th ed.; John Wiley & Sons, Inc.: New York, NY, USA, 1999; pp. 113–148.

59. Thompson, P.A. Compressible-Fluid Dynamics; McGraw-Hill: New York, NY, USA, 1972; pp. 102–103.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open accessarticle distributed under the terms and conditions of the Creative Commons Attribution(CC BY) license (http://creativecommons.org/licenses/by/4.0/).